Skip to main content
Log in

On the Stable Eigenvalues of Perturbed Anharmonic Oscillators in Dimension Two

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We study the asymptotic behavior of the spectrum of a quantum system which is a perturbation of a spherically symmetric anharmonic oscillator in dimension 2. We prove that a large part of its eigenvalues can be obtained by Bohr–Sommerfeld quantization rule applied to the normal form Hamiltonian and also admits an asymptotic expansion at infinity. The proof is based on the generalization to the present context of the normal form approach developed in Bambusi et al. (Commun Part Differ Equ 45:1–18, 2020) (see also Parnovski and Sobolev in Invent Math 181(3):467–540, 2010) for the particular case of \({\mathbb {T}}^d\).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Of course there are many technical details to verify, but this will be done in the forthcoming sections.

References

  1. Bambusi, D.: Exponential stability of breathers in Hamiltonian networks of weakly coupled oscillators. Nonlinearity 9(2), 433–457 (1996)

    Article  ADS  MathSciNet  Google Scholar 

  2. Bambusi, D., Berti, M., Magistrelli, E.: Degenerate KAM theory for partial differential equations. J. Differ. Equ. 250(8), 3379–3397 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  3. Bates, L.M., Fassò, F.: No monodromy in the champagne bottle, or singularities of a superintegrable system. J. Geom. Mech. 8(4), 375 (2016)

    Article  MathSciNet  Google Scholar 

  4. Bambusi, D., Fusè, A.: Nekhoroshev theorem for perturbations of the central motion. Regul. Chaotic Dyn. 22(1), 18–26 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  5. Bambusi, D., Fusè, A., Sansottera, M.: Exponential stability in the perturbed central force problem. Regul. Chaotic Dyn. 23(7–8), 821–841 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  6. Bambusi, D., Grebert, B., Maspero, A., Robert, D.: Reducibility of the quantum Harmonic oscillator in \(d\)-dimensions with polynomial time dependent perturbation. Anal. PDEs 11(3), 775–799 (2018)

    Article  MathSciNet  Google Scholar 

  7. Bambusi, D., Grebert, B., Maspero, A., Robert, D.: Growth of Sobolev norms for abstract linear Schrödinger Equations. J. Eur. Math. Soc. 23(2), 557–583 (2021)

    Article  MathSciNet  Google Scholar 

  8. Bambusi, D., Graffi, S., Paul, T.: Normal forms and quantization formulae. Commun. Math. Phys. 207(1), 173–195 (1999)

    Article  ADS  MathSciNet  Google Scholar 

  9. Bambusi, D., Langella, B., Montalto, R.: Growth of Sobolev norms for unbounded perturbations of the Laplacian on flat tori. Preprint, arXiv:2012.02654 (2020)

  10. Bambusi, D., Langella, B., Montalto, R.: On the spectrum of the Schrödinger operator on \({T}^d\): a normal form approach. Commun. Part. Differ. Equ. 45, 1–18 (2020)

    Article  Google Scholar 

  11. Bambusi, D., Langella, B., Montalto, R.: Spectral asymptotics of all the eigenvalues of Schrödinger operators on flat tori. Nonlinear Anal. 216, 112679 (2022)

    Article  Google Scholar 

  12. Charbonnel, A.-M.: Calcul fonctionnel a plusieurs variables pour des operateurs pseudodifferentiels dans \({R}^n\). Isr. J. Math. 45(1), 69–89 (1983)

    Article  MathSciNet  Google Scholar 

  13. Charbonnel, A.-M.: Spectre conjoint d’opérateurs pseudodifférentiels qui commutent. Ann. Fac. Sci. Toulouse Math. (5) 5(2), 109–147 (1983)

    Article  MathSciNet  Google Scholar 

  14. Charbonnel, A.-M.: Localisation et developpment asymptotique des elements du spectre conjoint d’operateurs pseudodifferentiels qui commutent. Integral Equ. Oper. Theory 9(4), 502–536 (1986)

    Article  MathSciNet  Google Scholar 

  15. de Verdière, Y.C.: Spectre conjoint d’opérateurs pseudo-différentiels qui commutent. II. Le cas intégrable. Math. Z. 171(1), 51–73 (1980)

    Article  MathSciNet  Google Scholar 

  16. Duistermaat, J.J.: On global action-angle coordinates. Commun. Pure Appl. Math. 33(6), 687–706 (1980)

    Article  MathSciNet  Google Scholar 

  17. Féjoz, J., Kaczmarek, L.: Sur le théorème de Bertrand (d’après Michael Herman). Ergod. Theory Dyn. Syst. 24(5), 1583–1589 (2004)

    Article  Google Scholar 

  18. Feldman, J., Knörrer, H., Trubowitz, E.: The perturbatively stable spectrum of a periodic Schrödinger operator. Invent. Math. 100(2), 259–300 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  19. Feldman, J., Knörrer, H., Trubowitz, E.: Perturbatively unstable eigenvalues of a periodic Schrödinger operator. Comment. Math. Helv. 66(4), 557–579 (1991)

    Article  MathSciNet  Google Scholar 

  20. Friedlander, L.: On the spectrum of the periodic problem for the Schrödinger operator. Commun. Part. Differ. Equ. 15(11), 1631–1647 (1990)

    Article  Google Scholar 

  21. Helffer, B., Robert, D.: Asymptotique des niveaux d’énergie pour des hamiltoniens à un degré de liberté. Duke Math. J. 49(4), 853–868 (1982)

    Article  MathSciNet  Google Scholar 

  22. Helffer, B., Robert, D.: Propriétés asymptotiques du spectre d’opérateurs pseudodifférentiels sur \({ R}^{n}\). Commun. Part. Differ. Equ. 7(7), 795–882 (1982)

    Article  Google Scholar 

  23. Hitrik, M., Sjöstrand, J., Vu Ngoc, S.: Diophantine tori and spectral asymptotics for nonselfadjoint operators. Am. J. Math. 129(1), 105–182 (2007)

    Article  MathSciNet  Google Scholar 

  24. Karpeshina, Y.E.: Perturbation series for the Schrödinger operator with a periodic potential near planes of diffraction. Commun. Anal. Geom. 4(3), 339–413 (1996)

    Article  Google Scholar 

  25. Parnovski, L.: Bethe-Sommerfeld conjecture. Ann. Henri Poincaré 9(3), 457–508 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  26. Popov, G.: Invariant tori, effective stability, and quasimodes with exponentially small error terms. I. Birkhoff normal forms. Ann. Henri Poincaré 1(2), 223–248 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  27. Popov, G.: Invariant tori, effective stability, and quasimodes with exponentially small error terms. II. Quantum Birkhoff normal forms. Ann. Henri Poincaré 1(2), 249–279 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  28. Parnovski, L., Sobolev, A.V.: Bethe-Sommerfeld conjecture for periodic operators with strong perturbations. Invent. Math. 181(3), 467–540 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  29. Parnovski, L., Shterenberg, R.: Complete asymptotic expansion of the integrated density of states of multidimensional almost-periodic Schrödinger operators. Ann. Math. (2) 176(2), 1039–1096 (2012)

    Article  MathSciNet  Google Scholar 

  30. Roy, N.: A semi-classical K. A. M. theorem. Commun. Part. Differ. Equ. 32(4–6), 745–770 (2007)

    Article  MathSciNet  Google Scholar 

  31. Rüssmann, H.: Invariant tori in non-degenerate nearly integrable Hamiltonian systems. Regul. Chaotic Dyn. 6(2), 119–204 (2001)

    Article  MathSciNet  Google Scholar 

  32. Sjöstrand, J.: Semi-exited states in nondegenerate potential wells. Asymptot. Anal. 6, 29–43 (1992)

    MATH  Google Scholar 

  33. Veliev, O.: Multidimensional Periodic Schrödinger Operator. Springer Tracts in Modern Physics. Perturbation Theory and Applications, vol. 263. Springer, Cham (2015)

    Google Scholar 

Download references

Acknowledgements

First of all we thank Didier Robert for pointing to our attention the papers by Charbonnel, then we thank Francesco Fassò for pointing to our attention the paper [BF16] and San Vũ Ngọc for several suggestions on quantum action angle coordinates. During the preparation of this work we also had several discussions with Alberto Maspero, that we warmly thank.

We thank the Italian Gruppo Nazionale di Fisica Matematica of INDAM for the support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Beatrice Langella.

Additional information

Communicated by C. Liverani.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Proof of Lemma 4.5

We start by proving the following easy lemma

Lemma A.1

The function \(a_r\) defined by (4.4) is analytic on the domain

$$\begin{aligned} 0<\left| L\right| <\left( \frac{2\ell }{\ell +1}E\right) ^{\frac{\ell +1}{2\ell }},\quad E>0; \end{aligned}$$
(A.1)

Proof

The effective Hamiltonian \(h_0^*(r,p_r,L)\) defined by (4.3) is an analytic function except at \(r=0\). Furthermore, for \(L\not =0\), \(h^*\), as a function of \(r,p_r\) is a submersion, except on the level surface of level \(E=|L|^{\frac{\ell +1}{2\ell }}\). So, outside this domain the level surface depend analytically on both E and L. Since \(a_r\) is just the normalized area contained in the level surface, it also depends analytically on EL in the considered domain. \(\square \)

We now study the behavior of \(a_r\) as \(L\rightarrow 0\). To this end we apply the method of the residues to the integral defining it. Precisely, we prove the following Lemma

Lemma A.2

Define

$$\begin{aligned} \rho := \frac{{L}^2}{E^{\frac{(\ell +1)}{\ell }}}, \end{aligned}$$
(A.2)

then there exists \(\rho _*\) and a function f(EL), analytic in the domain

$$\begin{aligned} E>0, \quad 0\le \rho <\rho _*, \end{aligned}$$
(A.3)

s.t.

$$\begin{aligned} a_r = -\frac{1}{2} |L| + f(E, L). \end{aligned}$$
(A.4)

Proof

Performing the change of variables \(r^2 =s,\) in the integral defining \(a_r\) one has

$$\begin{aligned} {a_r}= \frac{\sqrt{2}}{2 \pi } \int _{s_m}^{s_M} \frac{1}{s}\sqrt{-p(s, E, {L})}\ d s, \end{aligned}$$
(A.5)

with

$$\begin{aligned} {p(s, E, {L}) = \frac{s^{\ell +1}}{2\ell } - Es + \frac{{L}^2}{2}} \end{aligned}$$
(A.6)

and \(s_m = s_m(E, {L}),\) \(s_M=s_M(E, {L})\) the two positive solutions of the equation \({p(s, E, {L}) = 0.}\)

We now study the zeroes of \(p(s, E, {L})\). To this aim we observe that, since p is a polynomial in s with coefficients depending on E and \({L}\), it has \(\ell +1\) complex roots which of course depend continuously on E and L. Actually there is more structure: indeed \(s_*(E,L)\) is a root of p if and only if \(t_*=s_*E^{-1/\ell }\) is a root of

$$\begin{aligned} {{\tilde{p}}}(t) = \frac{t^{\ell +1}}{2\ell } - t + \frac{1}{2}\rho , \end{aligned}$$
(A.7)

which is a function of \(\rho \) only.

Since we are interested in a neighborhood of \(\rho =0\), we start by remarking that, for \(\rho =0\) the roots of \({{\tilde{p}}}\) are

$$\begin{aligned} t_0(0) = 0, \quad \left( t_j(0)\right) ^{\ell } = 2 \ell \quad \forall j = 1, \dots , \ell . \end{aligned}$$
(A.8)

To be determined we take \(t_1(0):=(2\ell )^{1/\ell }\) to be the positive real root of \(2\ell \) and the other roots in counterclockwise order. Correspondingly we will have

$$\begin{aligned} E^{1/\ell }t_0(\rho )=s_m(E,L),\quad E^{1/\ell }t_1(\rho )=s_M(E,L). \end{aligned}$$

Denote

$$\begin{aligned} d:=\min _{j\not =l}\left| t_j(0)-t_l(0)\right| , \end{aligned}$$

then there exists \(\rho _*\) s.t., for \(\rho <\rho _*\) one has

$$\begin{aligned} \min _{j\not =l}\left| t_j(\rho )-t_l(\rho )\right| \ge d/2 \end{aligned}$$

and correspondingly

$$\begin{aligned} \min _{j\not =l}\left| s_j(E, {L}) - s_{l}(E, {L})\right| \ge \frac{d}{2} E^{\frac{1}{\ell }} \end{aligned}$$
(A.9)

The function to be integrated in (A.5) is

$$\begin{aligned} F(z) := \frac{\sqrt{2}}{2\pi }\frac{1}{z} \sqrt{- p(z, E, {L})}, \end{aligned}$$
(A.10)

In order to make it holomorphic we cut \({\mathbb {C}}\) along the segments \(b_j\) joining \(s_{2(j-1)}(E, {L})\) with \(s_{2 j - 1}(E, {L})\), \(j=1,\ldots ,\lceil \frac{\ell }{2} \rceil \); if \(\ell \) is odd, there is a last cut \(b_{\lceil \frac{\ell }{2} \rceil }\), which is the half-line parallel to the real axis joining \(s_{\ell }(E, {L})\) with \(\infty \). Remark that \(b_1\) is the interval of integration in which we are interested.

We are now ready to choose the curve over we integrate to apply the method of the residue. To this end we define

$$\begin{aligned} \mathtt{M} := \max _{j=1,..,\ell , \rho <\rho _*} {\text {Re}} \left( t_j(\rho )\right) , \end{aligned}$$
(A.11)

and take the curve \(\Gamma _\varepsilon \) described in Fig. 1, with

$$\begin{aligned} \begin{aligned} a = \left( -\frac{d}{4}- \mathrm{i}\frac{d}{4}\right) E^{\frac{1}{\ell }}, \quad b= \left( 2\mathtt{M} - \mathrm{i}\frac{d}{4} \right) E^{\frac{1}{\ell }},\\ c =\left( 2 \mathtt{M} + \mathrm{i}\frac{d}{4}\right) E^{\frac{1}{\ell }}, \quad d = \left( -\frac{d}{4} + \mathrm{i}\frac{d}{4}\right) E^{\frac{1}{\ell }}. \end{aligned} \end{aligned}$$
(A.12)

We also denote \(\gamma _R\) the boundary of the rectangle abcd. With this notation we have

$$\begin{aligned} \begin{aligned} 2 a_r&= \lim _{\varepsilon \rightarrow 0} \int _{\Gamma _\varepsilon } F(z)\ d z - \int _{\gamma _R} F(z)\ d z\\&= 2 \pi \mathrm{i}\text {Res}(F, 0) - \int _{\gamma _R} F(z)\ d z=-|{L}| - \int _{\gamma _R} F(z)\ d z. \end{aligned} \end{aligned}$$
(A.13)

To get the thesis just define

$$\begin{aligned} f(E,L):= - \int _{\gamma _R} F(z)\ d z, \end{aligned}$$

and remark that this is analytic in the considered domain. \(\square \)

Fig. 1
figure 1

Representation of the path \(\Gamma _\varepsilon \). The radius of the circles is \(\varepsilon \), and the points abcd are defined as in (A.12)

Joining the results of these two lemmas one gets

Corollary A.3

There exists function \(f=f(E,L)\) analytic in the domain

$$\begin{aligned} \left\{ (E, L)\ \left| \ E>0,\ |L| < \left( \frac{2\ell }{\ell + 1} E\right) ^{\frac{\ell +1}{2 \ell }}\right. \right\} , \end{aligned}$$
(A.14)

such that

$$\begin{aligned} a_r(E, L) = -\frac{1}{2}|L| + f(E, L). \end{aligned}$$
(A.15)

We finally prove Lemma 4.5:

Proof of Lemma 4.5

Let

$$\begin{aligned} a_1 := {\left\{ \begin{array}{ll} a_r &{} \text {if } L\ge 0\\ a_r - L &{}\text {if } L< 0. \end{array}\right. } \end{aligned}$$
(A.16)

then, by Corollary A.15, there exists a function f analytic in the domain (A.14) s.t.

$$\begin{aligned} a_1=-\frac{L}{2}+f(E,L), \end{aligned}$$

so that \(a_1\) itself is analytic in such a domain. This concludes the proof of Items (1) and (2).

We come to the homogeneity properties. Take \(L\not =0\), then by (4.4) and performing the change of variables \(r = |L|^{\frac{1}{\ell +1}} x\) in (4.4), one sees that

$$\begin{aligned} a_r = \frac{\sqrt{2 E}}{\pi } |L|^{\frac{1}{\ell +1}} F(\rho ), \quad F(\rho ) = \int _{x_m}^{x_M} \sqrt{1-\rho \left( \frac{1}{2x^2} + \frac{x^{2\ell }}{2\ell }\right) }\,d x, \end{aligned}$$

with \(\rho \) as in (A.7) and \(x_m, x_M\) solving the equation \( \displaystyle {\rho \frac{x^{2\ell }}{2\ell } + \rho \frac{1}{2x^2} -1 = 0.}\) As a consequence, for any \(\lambda >0\) \(a_r\) satisfies

$$\begin{aligned} a_r(\lambda ^{\frac{2\ell }{\ell +1}}E, \lambda L) = \lambda ^{\frac{\ell }{\ell +1}} \sqrt{2 E} \lambda ^{\frac{1}{\ell +1}} |L| F(\rho ) =\lambda a_r(E, L), \end{aligned}$$

also \(a_1\) defined as in (A.16) satisfies

$$\begin{aligned} a_1(\lambda ^{\frac{2\ell }{\ell +1}}E, \lambda L) = \lambda a_r(E, L). \end{aligned}$$
(A.17)

Since \(h_0\) and \(a_2\) as in (4.2) are quasi-homogeneous functions of \(x, \xi \), one immediately deduces the quasi-homogeneity of \(a_1\). The homogeneity of \(h_0\) as a function of a also immediately follows from (A.17). We still have to consider the case \(L=0\). In this case the result immediately follows by continuity, by taking the limit of the functions to this set. \(\square \)

Appendix B: Proof of Lemma 5.12

First we prove the following

Lemma B.1

Let \(\Pi ^*:=a^{-1}(\mathop {\Pi }^\circ )\) and suppose \(f\in C^{\infty }({\mathbb {R}}^2)\) is such that \(f\circ \phi ^{\varphi }_a=f\), \(\forall \varphi \) then there exists \({{\tilde{f}}}\in C^{\infty }(\mathop \Pi ^{\circ })\) s.t. \(f={{\tilde{f}}}\circ a\) on \(\Pi ^*\).

Proof

Introducing action angle coordinates \((a,\varphi )\), which by the standard theory (see e.g. [Dui80]) are smooth and globally defined on the set \(\Pi ^*\), the function f is a \(C^\infty \) function of \((a,\varphi )\) which however does not depend on \(\varphi \). This is the wanted function \({{\tilde{f}}}\).

\(\square \)

Lemma B.2

There exist angle variables \(\varphi \) that are quasi-homogeneous of degree 0 as functions on \(\Pi ^*\), namely they satisfy

$$\begin{aligned} \varphi (\lambda x, \lambda ^\ell \xi ) = \varphi (x, \xi ) \quad \forall (x, \xi ) \in \Pi ^*,\ \forall \lambda \in {\mathbb {R}}^+. \end{aligned}$$

Proof

Let \({\varphi } = (\varphi _1, \varphi _2): \Pi ^* \rightarrow {\mathbb {T}}^2\) be such that \((a, {\varphi })\) are global action angle coordinates. For any choice of \({{\bar{a}}}\in \mathop \Pi ^{\circ }\) with \(\Vert {{\bar{a}}}\Vert =1\), let \((x_0, \xi _0) \in \Pi ^*\) be such that \(a(x_0,\xi _0)={{\bar{a}}}\) and \(\varphi (x_0, \xi _0) = 0\). Remark that it exists because the action \(\phi _a^{\varphi }\) on the level surfaces of a is transitive. Furthermore \((x_0,\xi _0)\) is a function of a only. For \((x, \xi ) \in \Pi ^*\), define \(\lambda \in {\mathbb {R}}^+\) and \(({{\widetilde{x}}},\widetilde{\xi })\) by

$$\begin{aligned} (x, \xi ) = (\lambda {\tilde{x}}, \lambda ^\ell {\tilde{\xi }}), \quad \text {and}\quad \Vert a({\widetilde{x}}, {\widetilde{\xi }})\Vert = 1. \end{aligned}$$
(B.1)

Observe that with this definition \(\lambda \) is a function of the actions a only. Then this implies that the function \({\tilde{\varphi }} = ({\tilde{\varphi }}_1, {\tilde{\varphi }}_2): \Pi ^* \rightarrow {\mathbb {T}}^2\) given by

$$\begin{aligned} {\tilde{\varphi }}(x, \xi ) := \varphi (x, \xi ) - \varphi (\lambda {x_0}, \lambda ^{\ell } {\xi _0}) \end{aligned}$$
(B.2)

still defines angle coordinates conjugated to the actions a on the set \(\Pi ^*\). This is due to the fact that \(\varphi (\lambda {x_0},\lambda ^\ell \xi _0)\) is a function of the actions only. Remark also that

$$\begin{aligned} {{\tilde{\varphi }}}(\lambda x_0,\lambda ^\ell \xi _0) =0,\quad \forall \lambda >0. \end{aligned}$$

From now on we use only the angles \({{\tilde{\varphi }}}\), so we omit the tildes.

We are now going to prove that these angles are homogeneous functions of degree 0 on \(\Pi ^*\).

To this aim, we define for \(\mu >0\) and \(j = 1,2\)

$$\begin{aligned} \varphi _{\mu , j}(x, \xi ):= \varphi _j(\mu x, \mu ^{\ell }\xi ) \end{aligned}$$

First of all, we observe that since \((a, \varphi )\) are canonically conjugated variables, one has \(\displaystyle {\left\{ a_j;\varphi _i\right\} = \delta _{i, j}}\), so one has

$$\begin{aligned} \begin{aligned} \left\{ a_i;\varphi _{\mu , j}\right\} (x, \xi ) = \mu ^{\ell }\partial _{x} a_i (x, \xi ) \cdot \partial _{\xi } \varphi _{j} (\mu x, \mu ^{\ell }\xi ) - \mu \partial _{\xi } a_i (x, \xi ) \cdot \partial _{x} \varphi _{j} (\mu x, \mu ^{\ell } \xi ). \end{aligned} \end{aligned}$$

Since \(a_i\) is quasi-homogeneous of degree \(\ell +1,\) \(\forall (x, \xi ) \in \Pi ^*\) one has

$$\begin{aligned} \begin{aligned} \mu ^{\ell }\partial _{x} a_i (x, \xi ) = \partial _{x} a_i(\mu x, \mu ^{\ell } \xi ), \\ \mu \partial _{\xi } a_i (x, \xi ) = \partial _{\xi }a_i(\mu x, \mu ^{\ell } \xi ), \end{aligned} \end{aligned}$$

one also obtains

$$\begin{aligned} \begin{aligned} \left\{ a_i;\varphi _{\mu , j}\right\} (x, \xi ) = \left\{ a_i;\varphi _{ j}\right\} (\mu x, \mu ^{\ell }\xi ) \equiv \delta _{i,j}. \end{aligned} \end{aligned}$$
(B.3)

Thus, using action angle coordinates to compute Poisson Brackets, one has

$$\begin{aligned} \left\{ a_j;\varphi _{\mu ,j}\right\} \equiv \frac{\partial \varphi _{\mu ,j}}{\partial \varphi _j} =1, \end{aligned}$$

therefore there exist functions \(f_{\mu , j}(a)\), depending on the actions only, such that

$$\begin{aligned} {\varphi }_{\mu , j} - {\varphi }_{j} = f_{\mu , j}(a). \end{aligned}$$

To prove that \(f_{\mu ,j}\) is identically zero, fix a value \({{\bar{a}}}\) of a, corresponding to some point \(({{\bar{x}}},{{\bar{\xi }}})\in \Pi ^*\) define \({\lambda }:=\Vert {{\bar{a}}}\Vert ^{\frac{1}{\ell +1}}\), then we have

$$\begin{aligned} f_{\mu , j}(a({{\bar{x}}},{{\bar{\xi }}}))&=f_{\mu , j}\left( \lambda ^{\ell +1}a\left( \frac{{{\bar{x}}}}{ \lambda },\frac{{{\bar{\xi }}}}{\lambda ^\ell }\right) \right) =f_{\mu , j}(\lambda ^{\ell +1}{a({x_0},{\xi _0})}) \\&=f_{\mu , j}\left( a\left( {\lambda } {{x_0}}, {\lambda ^\ell } {{\xi _0}}\right) \right) = {\varphi }_{\mu , j}\left( {\lambda } x_0, {\lambda ^\ell } \xi _0\right) - {\varphi }_{j} \left( {\lambda } x_0, {\lambda ^\ell } \xi _0\right) \\&=0-0=0. \end{aligned}$$

\(\square \)

Corollary B.3

Let \({\check{x}}, {\check{\xi }}: \Pi ^\circ \times {\mathbb {T}}^2\rightarrow \Pi ^* \) be the functions expressing the Cartesian coordinates in terms of the action angle coordinates, namely s.t.

$$\begin{aligned} ({\check{x}}, {\check{\xi }}) \circ (a, \varphi ) = \text {Id}, \quad (a, \varphi ) \circ ({\check{x}}, {\check{\xi }}) = \text {Id}. \end{aligned}$$

Then the following holds:

$$\begin{aligned} {\check{x}} (\lambda ^{\ell +1}a, \varphi ) = \lambda {\check{x}}(a, \varphi ), \quad {\check{\xi }} (\lambda ^{\ell +1}a, \varphi ) = \lambda ^{\ell } {\check{\xi }}(a, \varphi ) \quad \forall \lambda \in {\mathbb {R}}^+,\ (a, \varphi ) \in \Pi ^\circ \times {\mathbb {T}}^2.\nonumber \\ \end{aligned}$$
(B.4)

We now prove Lemma 5.12:

Proof

By Lemma B.1, there exists \({\tilde{f}} \in C^\infty (\mathop \Pi ^{\circ })\) such that \(f = {\tilde{f}} \circ a\). Passing to action angle variables on the set \(\Pi ^*\), one has

$$\begin{aligned} {\tilde{f}}(a) ={f} ({\check{x}}(a, \varphi ), {\check{\xi }}(a, \varphi )) \end{aligned}$$

First we prove the estimates for \(a\in \mathop \Pi ^{\circ }\). Consider for example

$$\begin{aligned} \partial _{a_1} {\tilde{f}} = \partial _x {f} \cdot \partial _{a_1} {\check{x}} + \partial _\xi {f} \cdot \partial _{a_1} {{\check{\xi }}} \end{aligned}$$

thus, using \({f} \in S_{AN,\delta }^{m},\) and the estimates

$$\begin{aligned} \left| \partial _a^{\beta }{\check{x}}(a, \varphi )\right| \lesssim _{} \langle a \rangle ^{\frac{1}{\ell + 1} - |\beta |}, \quad \left| \partial _a^{\beta }{{\check{\xi }}}(a, \varphi )\right| \lesssim _{} \langle a \rangle ^{\frac{1}{\ell + 1} - |\beta |}. \end{aligned}$$

which follow from (B.4) we can estimate such a quantity. One gets for \(a \in \Pi ^\circ \)

$$\begin{aligned} | \partial _a {\tilde{f}}| \lesssim \langle a \rangle ^{m-\delta _1 -\frac{\ell }{\ell + 1}} + \langle a \rangle ^{m-\delta _2 -\frac{1}{\ell + 1}}\simeq \langle a\rangle ^{m-\varsigma }. \end{aligned}$$

Iterating and studying the other derivatives, one also has that \(\forall \alpha \in {\mathbb {N}}^2\), implies

$$\begin{aligned} | \partial _a^\alpha {\tilde{f}}| \lesssim \langle a \rangle ^{m-\varsigma |\alpha |}, \end{aligned}$$

but just for \(a \in \Pi ^\circ \). To get a symbol defined on the whole of \({\mathbb {R}}^2\), we consider again the cones \({{\mathcal {V}}}\) and \({{\mathcal {C}}}\) and the cutoff function \(\Psi \) supported in \({{\mathcal {V}}}\) and equal to one in \({{\mathcal {C}}}\), and just consider the function

$$\begin{aligned} f_c(a):={{\tilde{f}}}(a)\Psi (a)(1-\chi (\left\| a\right\| )), \end{aligned}$$

which has all the claimed properties. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bambusi, D., Langella, B. & Rouveyrol, M. On the Stable Eigenvalues of Perturbed Anharmonic Oscillators in Dimension Two. Commun. Math. Phys. 390, 309–348 (2022). https://doi.org/10.1007/s00220-021-04301-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-021-04301-w

Navigation